Free Electrons: 7.1 Plasma Re Ectivity
Free Electrons: 7.1 Plasma Re Ectivity
7
In this chapter we investigate the optical properties associated with free
electrons. As the name suggests, these are systems in which the electrons
7.1 Plasma reflectivity 180 experience no restoring force from the medium when driven by the elec-
7.2 Free carrier tric field of a light wave. The two main solid-state systems that exhibit
conductivity 183
strong free electron effects are:
7.3 Metals 185
7.4 Doped • Metals. Metals contain large densities of free electrons that orig-
semiconductors 191 inate from the valence electrons of the metal atoms.
7.5 Plasmons 198
• Doped semiconductors. n-type semiconductors contain free elec-
7.6 Negative refraction 207 trons, while p-type materials contain free holes. The free carrier
Chapter summary 209 density is determined by the concentration of impurities used for
Further reading 210 the doping.
Exercises 211
We begin our discussion of the optical properties by using the Drude–
Lorentz model introduced in Section 2.1.3 of Chapter 2. This will enable
us to explain the main optical property of metals that we mentioned
in Section 1.4.3, namely that they reflect strongly in the visible spec-
tral region. We then apply our knowledge of interband transitions from
Chapter 3 to obtain a better understanding of the detailed form of the
reflectivity spectra of metals such as aluminium and copper. Next we ap-
ply the Drude–Lorentz model to doped semiconductors to explain why
doping causes infrared absorption. We then consider the collective oscil-
lations of the whole free carrier gas, which will naturally lead us to the
notion of plasmons, both bulk and surface. Finally, we conclude with a
brief discussion of negative refraction.
1.0
0.8
Reflectivity
0.6
0.4
0.2
0.0
Fig. 7.1 Reflectivity of an undamped 0 1 2
free carrier gas as a function of fre- w / wp
quency.
j = −N ev = σE , (7.12)
Example 7.2
The DC electrical conductivity of copper is 6.5 × 107 Ω−1 m−1 at room
temperature. Calculate the skin depth at 50 Hz and 100 MHz.
Solution
The skin depth is given by eqn 7.20. At a frequency of 50 Hz we have
ω = 2π × 50 = 314 rad/s. On inserting this value of ω into eqn 7.20
and using σ0 = 6.5 × 107 Ω−1 m−1 , we obtain δ = 8.8 mm. At 100 MHz,
ω = 6.28 × 108 rad/s, and the skin depth δ is only 6.2 µm.
7.3 Metals
The free electron model of metals was proposed by P. Drude in 1900. The
model provides a basic explanation for why metals are good conductors
of heat and electricity, and is the starting point for more sophisticated
theories. As we shall see here, it is also successful in explaining why
metals tend to be good reflectors. On the other hand, band theory is
needed to explain why some metals (e.g. copper and gold) are coloured.
1.0
Example 7.3
The conductivity of aluminium at room temperature is 4.1×107 Ω−1 m−1 .
Calculate the reflectivity at 500 nm according to the Drude–Lorentz
model.
Solution
We first work out the damping time τ from the conductivity using
eqn 7.14. Taking the value of N = 1.81 × 1029 m−3 from Table 7.1,
we find:
m0 σ 0
τ= = 8.0 × 10−15 s .
N e2
188 Free electrons
Table 7.1 also gives us the value of the plasma frequency, namely ωp =
2.4 × 1016 rad/s. The wavelength of 500 nm corresponds to an angular
frequency ω = 2πc/λ = 3.8 × 1015 rad/s. We use these frequencies in
eqns 7.16 and 7.17 to calculate the real and imaginary parts of the
complex dielectric constant:
ωp2 τ 2
1 = 1 − = −39 ,
1 + ω2 τ 2
and
ωp2 τ
2 = = 1.3 .
ω(1 + ω 2 τ 2 )
We then work out the real and imaginary parts of the complex refractive
index using eqns 1.25 and 1.26. This gives:
1 1/2
n = √ − 39 + [(−39)2 + (1.3)2 ]1/2 = 0.10 ,
2
and
1 1/2
κ = √ + 39 + [(−39)2 + (1.3)2 ]1/2 = 6.2 .
2
We finally obtain the reflectivity from eqn 1.29:
(n − 1)2 + κ2 (−0.9)2 + (6.2)2
R= = = 99% .
(n + 1)2 + κ2 (1.1)2 + (6.2)2
This shows that the inclusion of the damping reduces the reflectivity by
only 1% in this case.
Aluminium
The band diagram of aluminium is shown in Fig. 7.3. Aluminium has an
electronic configuration of [Ne]3s2 3p1 with three valence electrons. The
crystal structure is face-centred cubic, which has a body-centred cubic
7.3 Metals 189
16
Al
Energy (eV)
12 EF
8
Fig. 7.3 Band diagram of aluminium.
4 The transitions at the W and K points
that are responsible for the reflectivity
dip at 1.5 eV are labelled. After Segall
0
G X W L G K X (1961), c American Physical Society,
reprinted with permission.
Copper
Copper has an electronic configuration of [Ar]3d10 4s1 . The outer 4s
bands approximate reasonably well to free electron states with a dis-
persion given by E = 2 k 2 /2m0 . They therefore form a broad band
covering a wide range of energies. The 3d bands, on the other hand,
are more tightly bound and are relatively dispersionless, occupying only
a narrow range of energies. The density of states for the two bands is
illustrated schematically in Fig. 7.4. The narrow 3d bands can hold ten
electrons, and therefore their density of states is sharply peaked. The
4s band, which can hold two electrons, is much broader, with a smaller
maximum. The 11 valence electrons of copper fill up the 3d band, and
half fill the 4s band. The Fermi energy therefore lies within the 4s band
In atomic physics, transitions between above the 3d band. Interband transitions are possible from the filled
d and s states are forbidden for electric- 3d bands to unoccupied states in the 4s band above EF , as illustrated
dipole processes. (See Table B.1 in Ap- in Fig. 7.4. This implies that there will be a well-defined threshold for
pendix B.) The matrix element for
the 3d→4s transitions is therefore rel- interband transitions from the 3d bands to the 4s band.
atively small, but this is compensated Figure 7.5 shows the actual band structure and density of states of
by the very high density of states in copper. The general features indicated in Fig. 7.4 are clearly shown in
the solid, which results in strong ab- the calculated curves. The 4s band is the parabola starting at the Γ point
sorption.
at −9 eV, while the 3d bands are the five curves bunched together in the
energy range −5 → −2 eV. The 4s band crosses the 3d bands and then
re-emerges as the single band with energy > −2 eV. It is apparent that
the 3d electrons lie in relatively narrow bands with very high densities
E of states, while the 4s band is much broader with a lower density of
states. The Fermi energy lies in the middle of the 4s band above the
4s band 3d band. Interband transitions are possible from the 3d bands below
EF to unoccupied levels in the 4s band above EF . The lowest energy
optical transitions are marked on the band diagram in Fig. 7.5. The transition
transitions energy is 2.2 eV which corresponds to a wavelength of 560 nm.
EF Figure 7.6 shows the measured reflectivity of copper from the infrared
to the ultraviolet spectral region. Based on the plasma frequency given
in Table 7.1, we would expect near 100% reflectivity for photon energies
below 10.8 eV, which corresponds to an ultraviolet wavelength of 115 nm.
3d band
However, the experimental reflectivity falls off sharply above 2 eV owing
to the interband absorption edge discussed above. This explains why
Density of states
copper has a reddish colour.
Energy (eV)
-2 -2
3d
-4 bands -4
-6 -6
copper
-8 3d and 4s bands -8
4s band
W L G X W K 0 2 4 6
Density of states (states eV-1)
Wave vector k
Fig. 7.5 Calculated band structure of copper. The transitions from the 3d bands responsible for the interband transitions
around 2 eV are identified. The right-hand side of the figure shows the density of states calculated from the band structure. The
strongly peaked features between about −2 eV and −5 eV are due to the 3d bands. The dotted line is the integrated density of
states. The Fermi level (defined here as E = 0) corresponds to the energy at which the integrated density of states is equal to
11. After Moruzzi et al. (1978).
1.0
0.8 copper
Reflectivity
0.6
0.4
hwp
visible
0.2
Fig. 7.6 Reflectivity of copper from
the infrared to the ultraviolet spectral
region. The reflectivity drops sharply 0.0
above 2 eV due to interband transi-
0 2 4 6 8 10 12 14
tions. Data from Lide (1996). Energy (eV)
100
n-type InSb
80 4.0 ´ 1024
Reflectivity (%)
2.8 ´ 1024
1.2 ´ 1024
60 6.2 ´ 1023
3.5 ´ 1023
40
20
Fig. 7.7 Infrared reflectivity spectra of
n-type InSb at room temperature for
0 different values of the free electron den-
10 20 30 sity. After Spitzer & Fan (1957), c
Wavelength (mm) American Physical Society, reprinted
with permission.
By fitting this formula to the data, the effective mass of InSb can be
determined. (See Exercise 7.9.)
At frequencies above ωp , the presence of free carriers leads to the ab-
sorption of light. This effect is called free carrier absorption, and
can be observed in the infrared spectral region below the fundamental
absorption edge at the band gap, where the semiconductor would nor-
mally be transparent. To see how this effect arises, we split the dielectric
constant given in eqn 7.23 into its real and imaginary parts. This gives:
ωp2 τ 2
1 = opt 1 − , (7.26)
1 + ω2τ 2
opt ωp2 τ
2 = , (7.27)
ω(1 + ω 2 τ 2 )
E
hk 2
E= scattering
2me*
hw
EF
E
Fig. 7.9 Intervalence band absorption
in a p-type semiconductor. EF is the
Fermi energy determined by the dop-
k
D EF ing density. The labelled arrows indi-
cate: (1) transitions from the light-hole
hh band (lh) band to the heavy-hole (hh) band;
lh band (2) transitions from the split-off (SO)
SO band
band to the lh band; and (3) transitions
from the SO band to the hh band.
to conserve momentum.
1.0
Absorption (103 m-1) 2p±
2p0
3p±
Si : P
T = 4.2 K 4p±
3p0 5p± Fig. 7.11 Infrared absorption spec-
4p0 trum of n-type silicon doped with phos-
phorus at a density of 1.2 × 1020 m−3 .
0 The temperature was 4.2 K. After Ja-
32 34 36 38 40 42 44 gannath et al. (1981), c American
Photon energy (meV) Physical Society, reprinted with per-
mission.
electrons are excited from the top of the valence band to the empty
donor levels.
The valence band → donor level transitions occur at photon energies
In many direct-gap semiconductors it is
just below the band gap Eg , with a threshold given by Eg −E1D . However,
found experimentally that the absorp- the transitions tend to be broadened into a continuum both by thermal
tion decreases exponentially below the effects and by the fact that transitions can take place from a whole
band gap. This is called the Urbach tail range of states within the valence band. Hence the impurity transitions
on account of Urbach’s rule, which
states that the frequency dependence of cause a smearing of the absorption edge compared with the abrupt edge
the absorption for ~ω < Eg is given by: found at the band gap of pure semiconductors. The absorption strength
σ(~ω − E )
will always be weak compared to the interband and excitonic transitions
g
α(~ω) ∝ exp , due to the relatively small number of impurity atoms compared to the
kB T
where σ is a phenomenological fitting
density of states within the conduction band. On the other hand, the
parameter. transitions occur in the spectral region just below the band gap where
we would normally expect no absorption at all. Hence these transitions
do have an effect on the fundamental absorption edge, and make precise
determinations of Eg from the absorption spectra at room temperature
more difficult.
We have restricted our attention here to n-type semiconductors for
the sake of simplicity. The same effects can of course occur in p-type
materials.
7.5 Plasmons
Equation 7.7 tells us that the relative permittivity of a lightly damped
gas of free electrons is expected to be zero at ωp . This suggests that
something interesting might happen at this frequency. This is indeed
the case, as we discuss here.
∂j ∂2E 1 ∂B
+ 0 2 = ∇×
∂t ∂t µ0 ∂t
(7.37)
1
=− ∇× (∇ × E) .
µ0
The electrons will move in response to the local electric field according
to their equation of motion:
E = Et + El , (7.41)
It becomes apparent that both sides of This implies that we have two independent equations of motion for the
eqn 7.42 must be equal to zero by tak- transverse and longitudinal components:
ing the divergence and curl of the equa-
tion. ∂ 2E t
+ ωp2 E t − c2 ∇2 E t = 0, (7.43)
∂t2
∂2E l
+ ωp2 E l = 0. (7.44)
∂t2
We look for wave-like solutions in which the field varies with time and
position as exp i(k · r − ωt). For the transverse solutions we find:
c2 k 2 = ω 2 − ωp2 . (7.45)
ω = ωp . (7.46)
This shows that the medium can support longitudinal modes at the
In a metal, the frequency of the longi- plasma frequency, and that the modes are dispersionless: i.e. ω is in-
tudinal modes does in fact vary slightly dependent of k. These longitudinal modes correspond to the plasma
with k. The correction term is very oscillations that we discussed qualitatively at the start of the section.
small, and arises from the breakdown
Figure 7.12(a) shows a schematic diagram of the longitudinal electron
of some of the approximations used in
this derivation. See Exercise 7.18. displacements within a plasma oscillation and the fields that they gen-
erate.
The existence of longitudinal solutions at the plasma frequency is a
consequence of the fact that r = 0 at ωp . In a medium with zero average
7.5 Plasmons 201
charge density, Gauss’s law (eqn A.10) combined with eqn A.3 tells us
that
∇ · D = ∇·(r 0 E) = 0 . (7.47)
If r = 0, we then deduce that ∇ · E = 0. This is the normal situa-
tion for transverse electromagnetic waves in which the electric field is
perpendicular to the direction of the wave. However, if r = 0, we can
We shall come across another exam-
satisfy eqn 7.47 with waves that have ∇ · E = 0, i.e. longitudinal waves. ple of longitudinal modes at frequencies
We thus conclude that a dielectric can support longitudinal electric field where r = 0 when we consider phonons
waves at frequencies that satisfy r (ω) = 0. in Chapter 10. (See Section 10.2.2.)
Equation 7.44 shows us that the longitudinal oscillations of the plasma
behave as harmonic oscillators with a natural resonant frequency at ωp .
The derivation is completely classical, and the oscillator can have any
energy. However, we know in fact that the energy of harmonic oscillators
is quantized. We therefore expect the energy of the plasma oscillations
to be quantized in units of ωp . The quasi-particles that correspond to
these quantized plasma oscillations are called plasmons. As shown in
eqn 7.46, the frequency of the plasmons is independent of their wave
vector.
Since plasmons are associated with longitudinal plasma oscillations,
they cannot be excited directly by light, which is a transverse wave. In-
stead, they have to be observed by techniques of inelastic scattering, in
which a beam of particles excites plasmons while passing through the
medium, as illustrated in Fig. 7.12(b). The energy Ein of the incoming
particles must be significantly larger than the plasmon energy. Conser-
vation of energy requires that:
where Eout is the energy of the outgoing particle, and n is the number
of plasmons emitted. The detection of particles with energies given by
eqn 7.48 establishes that plasmons have been excited.
In the case of metals, the plasmon energies are several eV, and so
electrons with keV energies are typically used. By measuring the en-
ergy spectrum of the electrons emerging from a thin sample, the plasma
202 Free electrons
ωp2
m = 1 − . (7.59)
ω2
The relative permittivity of the dielectric is assumed to be real and
independent of frequency.
Consider first the dispersion in the bulk of the metal, which is shown in
Fig. 7.16(a). For frequencies below ωp , the light is reflected, and there are
7.5 Plasmons 205
w w
(a) (b)
w = ck w = ckx /Öòd
wp wp Fig. 7.16 (a) Photon dispersion in the
wsp bulk of a metal with a dielectric func-
tion given by eqn 7.7. (b) Surface plas-
mon polariton dispersion for the same
metal. The curve is drawn for the case
0 0 where the dielectric is air (i.e. d = 1).
0 1 2 0 1 2
ck /wp ckx /wp The dashed line in (b) shows the dis-
persion for light in the dielectric.
In region (1) at large kx the group velocity (i.e. dω/dk) is zero and
ω → ωsp . This asymptotic frequency limit is called the surface plasmon
frequency. Equation 7.58 shows us that kx → ∞ when (m +d ) → 0, and
so we can find ωsp by solving m (ω) = −d . For an undamped plasma
with m (ω) given by eqn 7.59, we find: In electron energy-loss experiments on
metals, it is common to observe peaks
ωp corresponding to both bulk and sur-
ωsp = √ . (7.61) face plasmons. This gives a convenient
1 + d
method for measuring both ωp and ωsp .
√
Note that ωsp = ωp / 2 when the dielectric is air.
The behaviour of the surface plasmon polaritons in region (1) where
ω < ωsp is the most interesting, since this corresponds to propagating
modes for frequencies below the plasma frequency. The spatial extent of
the fields in the z direction can be found from the values of kz in the
206 Free electrons
metal and the dielectric. This can be done by substituting eqn 7.58 back
into 7.57 to obtain:
1/2
d ω −2d
kz = ,
c m + d
1/2 (7.62)
m ω −2m
kz = .
c m + d
Since (m + d ) goes to zero at ωsp , the decay constants increase with
ω and diverge on approaching ωsp . The field decay length lz is equal
to 1/|kz |, and this implies that the polaritons become more localized
as ω approaches ωsp . Note that it is usually the case that |kzm | > |kzd |,
and hence that the plasmon extends further in the dielectric than in the
metal, as illustrated in Fig. 7.15. At optical frequencies, lzd and lzm are
typically a few hundred or few tens of nanometres, respectively. (See
Exercise 7.20.)
Two other branches of plasmonics deal One of the aims of the research field of plasmonics is to propagate
with the enhancement of radiative ef- electromagnetic waves in a metal as surface plasmon polaritons. With
ficiencies and the increased transmis- values of lzm being in the sub-100 nm range, the waves can be confined
sion of light through sub-wavelength
apertures. In the first case, the radia-
to dimensions smaller than the wavelength of light, and we then enter
tive efficiency of an atom in a dielec- the realm of nanophotonics, which is not accessible to conventional
tric can be enhanced by placing it close optics due to diffraction limits. The distance that the plasmon modes
to a metal surface, and hence exploit- can propagate along the surface is determined by the imaginary part of
ing the large electric field amplitude
that is present at the interface. In the m . This can be several tens of micrometres (see Exercise 7.20), which
second case, it has been demonstrated is more than adequate for the applications that are being considered.
that plasmonic effects can enhance the An issue that has to be addressed in plasmonics is the way to couple
transmission of light through periodic light to the polaritons. It is apparent from Fig. 7.16(b) that the polariton
arrays of sub-wavelength-size holes in
an optically thick metallic film. Details modes always lie below the light line. This means that polaritons can
of these and other applications may be never transform directly into light: they are non-radiative modes. By the
found in the Further Reading. same token, it is not possible to couple light directly from the dielectric
into the polariton modes, since it is never possible to match the wave
vectors. Therefore, in order to couple external light into the surface
plasmon polaritons, techniques must be used to change the wave vector
of the light. One way to do this is to use a grating. In fact, it has been
known for a long time that the reflectivity of metallic ruled gratings can
drop significantly when one of the diffracted orders propagates parallel
to the surface. This effect, which is called Wood’s anomaly, is now
known to be caused by the excitation of surface plasmon polaritons in
the metal. Details of how the coupling to polariton modes is achieved in
practice may be found in the references cited for Further Reading.
A striking example of surface plasmon effects is to be found in consid-
Colloids can be solids, liquids, or gases.
The particle size in a colloid should
ering the optical properties of metal colloids. Metal colloids are made by
be smaller than the wavelength of dispersing a large number of very small metallic particles (i.e. ‘nanopar-
light. An interesting application of ticles’) throughout a homogeneous medium such as glass or water. It
metal colloids is in the production of has been known from antiquity that the colour of metal colloids is dif-
stained glass. The colouration of me-
dieval stained glass is typically caused
ferent from that of the bulk metal, and this phenomenon is now known
by gold, silver or copper nanoparticles to be caused by the resonant excitation of surface plasmons in the metal
incorporated into the glass during the nanoparticles. However, in contrast to surface plasmon polaritons, the
melt process. surface fields in the metallic nanoparticles do not propagate, since they
7.6 Negative refraction 207
1.2
Chapter summary
Further reading
The properties of electromagnetic waves in a conducting sorption.
medium are covered in many electromagnetism and op- The properties of plasmons are treated in depth in
tics textbooks, for example Bleaney and Bleaney (1976), Maier (2007). Bulk plasmons are covered in Kittel (2005),
Born and Wolf (1999), or Hecht (2001). while the classic text on surface plasmons is Raether
The free carrier model of metals is covered in Singleton (1988). Review articles on the research fields of plas-
(2001). It is also covered in Ashcroft and Mermin (1976), monics and nanophotonics may be found in Barnes et
Burns (1985), or Kittel (2005). al. (2003), Lal et al. (2007), Maier and Atwater (2005),
Free carrier reflectivity and absorption in semiconduc- Murray and Barnes (2007), or Ebbeson et al. (2008).
tors has been reviewed by Pidgeon (1980), and is also The concept of negative refraction was proposed in
covered by Yu and Cardona (1996). Yu and Cardona give Veselago (1968). Introductory reviews on the subject may
further details about intervalence band and impurity ab- be found in Pendry (2004) or Pendry & Smith (2004).
Exercises 211
A more detailed review is given in Ramakrishna (2005). quencies is reviewed in Shalaev (2007).
Progress on obtaining negative refraction at optical fre-
Exercises
(7.1) Derive a relationship between the Fermi energy EF (7.6) Estimate the fraction of light with wavelength
of a metal and its plasma frequency ωp . 1 m that is transmitted through a 20 nm thick
(7.2) The ionosphere reflects radio waves with frequen- gold film at 77 K, where the DC electrical conduc-
cies up to about 3 MHz, but transmits waves with tivity is 2 × 108 Ω−1 m−1 . The plasma frequency
higher frequencies. Estimate the free electron den- and electron density of gold are given in Table 7.1.
sity in the ionosphere. (7.7) Figure 7.20 shows the measured reflectivity of gold
(7.3) Estimate the skin depth of radio waves of fre- in the wavelength range 100–1000 nm. Account
quency 200 kHz in sea water, which has an av- qualitatively for the shape of the spectrum, and
erage electrical conductivity of about 4 Ω−1 m−1 . deduce the energy gap between the d bands and
Hence discuss the difficulties you might encounter the Fermi energy. Use the data to explain the char-
when attempting to communicate with a sub- acteristic colour of gold.
merged submarine using radio waves. (7.8) What is the value of the dielectric constant of a
(7.4) Cesium metal is found to be transparent to electro- medium that has zero reflectivity? Use eqn 7.22
magnetic radiation of wavelengths below 440 nm. to show that the reflectivity of a lightly damped
Calculate a value for the electron effective mass doped semiconductor is zero at the angular fre-
using the data given in Table 7.1. quency given in eqn 7.25.
(7.5) The momentum scattering time of silver is 4.0 × (7.9) Use the data shown in Fig. 7.7 to deduce the value
10−14 s at room temperature. Calculate the dielec- of the electron effective mass of InSb at each car-
tric constant at 500 nm, neglecting interband ab- rier density. Take opt = 15.6.
sorption effects. Hence estimate the reflectivity of (7.10) The absorption coefficient at room temperature of
a silver mirror at this wavelength. See Table 7.1 an n-type sample of InAs with a doping level of
for the plasma frequency of silver. 1.4×1023 m−3 is found to be 500 m−1 at 10 m. Es-
timate the momentum scattering time, given that
the electron effective mass is 0.023 m0 and the re-
1.0 fractive index is 3.5.
gold (7.11)∗ A laser beam operating at 632.8 nm with an inten-
0.8 sity of 106 W m−2 is incident on a sample of pure
Reflectivity
The valence band parameters for GaAs are given (7.18) In a metal, the frequency at which r = 0 varies
in Table D.2. slightly with the wave vector k ≡ |k|, and for small
(a) Work out the Fermi energy in the valence band k we have:
on the assumption that the holes are degenerate. 3vF2 k 2
What are the wave vectors of the heavy and light ω(k) ≈ ωp 1 + ,
10ωp2
holes at the Fermi energy?
(b) Calculate the upper and lower limits of the where vF is the Fermi velocity of the electrons.
photon energies for the three absorption processes Consider a metal with electron density N =
labelled (1), (2), and (3) in Fig. 7.9, namely the 1029 m−3 and lattice constant a = 4 Å. Calculate
lh → hh, the SO → lh and the SO → hh transi- the relative size of the term in k2 for k = 0.1 π/a,
tions. i.e. 10% of the size of the Brillouin zone.
(7.13) Figure 7.11 shows the infrared absorption spec- (7.19) In an electron energy-loss experiment on a metal,
trum of n-type silicon, which has a dielectric con- two series of peaks are observed that obey eqn 7.48
stant of 16. Two series of lines labelled np0 and with ~ωp = 10.3 eV and 15.3 eV. Use the data in
np± are identified in the data. Table 7.1 to determine which metal is being inves-
(a) Show that the np0 series is consistent with tigated, and account for the two series of peaks.
eqn 7.30, and deduce a value for the electron ef- (7.20) A surface plasmon mode with a frequency corre-
fective mass for these transitions. sponding to a vacuum wavelength of 600 nm is ex-
(b) Show that the np± series follows the following cited at the interface between silver and air. The
formula: relative permittivity of silver at this frequency is
R∗ R∗
hν = 20 − ± , given approximately by m = −18 + i.
1 n2 (a) Calculate the decay constants kzd and kzm , and
∗ ∗
stating the values of R0 and R± deduced from the hence deduce the field decay lengths in the direc-
data. tion normal to the surface in both the air and in
(7.14) It is found that the infrared absorption spectrum the metal.
of a lightly doped n-type semiconductor with a (b) The propagation length in the direction par-
relative permittivity of 15.2 consists of a series of allel to the surface is defined as the distance over
sharp lines at low temperatures. The energies of which the intensity drops by a factor of e−1 . Cal-
the lines are given by: culate the imaginary part of kx , and hence deduce
the propagation length for silver at 600 nm.
E(n) = R∗ (1 − 1/n2 )
(7.21) Show that the resonance of the polarizabil-
where R∗ is 2.1 meV and n is an integer greater ity of colloidal metal nanoparticles that obey
than 1. Explain why the energies of the absorp- eqn √
7.7 occurs at an angular frequency given by
tion lines are almost independent of the type of ωp / 1 + 2n2 in a medium with a refractive index
impurity atoms used for the doping, and deduce a of n. Evaluate this frequency for the case where
value for the electron effective mass. the dielectric is air.
(7.15) The fundamental absorption edge of a semicon- Table 7.3 Complex relative permittivity of gold be-
ductor shifts from 5.26 m to 5.44 m when doped tween 500 and 550 nm. Adapted from Raether (1988).
with acceptors. Deduce a value for the ground
state acceptor level energy relative to the valence Wavelength (nm) 1 2
band.
500 –2.3 3.6
(7.16) The beam from an argon ion laser operating at 510 –3.0 3.1
514.5 nm is incident on an n-type GaAs sample. 520 –3.7 2.7
A peak is observed in the intensity of the scat- 530 –4.4 2.4
tered light at 534.3 nm. Explain this observation, 540 –5.2 2.2
550 –6.0 2.0
and estimate the electron density, given that m∗e =
0.067 m0 and n = 3.3.
(7.17) Calculate the doping density at which the plas- (7.22) (a) The complex relative permittivity of gold in the
mons in n-type GaAs have the same wave number range 500–550 nm is given in Table 7.3. Use this
as the longitudinal optic phonon at 297 cm−1 . Take information to estimate the resonance wavelength
m∗e = 0.067 m0 and n = 3.3. of colloidal gold nanoparticles in water, which has
Exercises 213
a refractive index of 1.33. diction of the previous exercise. Account for any
(b) Use the data for gold given in Table 7.1 to com- difference.
pare the result obtained in part (a) with the pre-